You are on page 1of 15

ll

OPEN ACCESS

Article
SARS-CoV-2 Disrupts Splicing, Translation,
and Protein Trafficking to Suppress Host Defenses
Abhik K. Banerjee,1,2,8 Mario R. Blanco,1,8 Emily A. Bruce,3,8 Drew D. Honson,1,9 Linlin M. Chen,1,9 Amy Chow,1,9
Prashant Bhat,1,4 Noah Ollikainen,1 Sofia A. Quinodoz,1 Colin Loney,5 Jasmine Thai,1 Zachary D. Miller,6 Aaron E. Lin,7
Madaline M. Schmidt,3 Douglas G. Stewart,5 Daniel Goldfarb,5 Giuditta De Lorenzo,5 Suzannah J. Rihn,5
Rebecca M. Voorhees,1 Jason W. Botten,3 Devdoot Majumdar,6,* and Mitchell Guttman1,10,*
1Division of Biology and Biological Engineering, California Institute of Technology, Pasadena, CA 91125, USA
2Keck School of Medicine, University of Southern California, Los Angeles, CA 90089, USA
3Departments of Medicine, Division of Immunobiology and Microbiology, and Molecular Genetics, Larner College of Medicine, University of

Vermont, Burlington, VT 05405, USA


4David Geffen School of Medicine, University of California, Los Angeles, Los Angeles, CA 90095, USA
5MRC-University of Glasgow Centre for Virus Research (CVR), Glasgow G61 1QH, UK
6Department of Surgery and University of Vermont Cancer Center, University of Vermont College of Medicine, 89 Beaumont Avenue,

Burlington, VT 05405, USA


7Broad Institute of MIT and Harvard, Cambridge, MA 02142, USA
8These authors contributed equally
9These authors contributed equally
10Lead Contact

*Correspondence: dev.Majumdar@med.uvm.edu (D.M.), mguttman@caltech.edu (M.G.)


https://doi.org/10.1016/j.cell.2020.10.004

SUMMARY

Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is a recently identified coronavirus that
causes the respiratory disease known as coronavirus disease 2019 (COVID-19). Despite the urgent need,
we still do not fully understand the molecular basis of SARS-CoV-2 pathogenesis. Here, we comprehensively
define the interactions between SARS-CoV-2 proteins and human RNAs. NSP16 binds to the mRNA recog-
nition domains of the U1 and U2 splicing RNAs and acts to suppress global mRNA splicing upon SARS-CoV-2
infection. NSP1 binds to 18S ribosomal RNA in the mRNA entry channel of the ribosome and leads to global
inhibition of mRNA translation upon infection. Finally, NSP8 and NSP9 bind to the 7SL RNA in the signal
recognition particle and interfere with protein trafficking to the cell membrane upon infection. Disruption of
each of these essential cellular functions acts to suppress the interferon response to viral infection. Our re-
sults uncover a multipronged strategy utilized by SARS-CoV-2 to antagonize essential cellular processes to
suppress host defenses.

INTRODUCTION 2020). These include 4 structural proteins: the nucleocapsid (N;


which binds the viral RNA) and the envelope (E), membrane (M),
Coronaviruses are a family of viruses with notably large single- and spike (S) proteins, which are integral membrane proteins. In
stranded RNA genomes and broad species tropism among mam- addition, there are 16 non-structural proteins (NSP1–NSP16)
mals (Graham and Baric, 2010). Recently, a coronavirus, severe that encode the RNA-directed RNA polymerase, helicase, and
acute respiratory syndrome coronavirus 2 (SARS-CoV-2), was other components required for virus replication (da Silva et al.,
discovered to cause the severe respiratory disease known as co- 2020). Finally, there are 7 accessory proteins (ORF3a–ORF8)
ronavirus disease 2019 (COVID-19). It is highly transmissible in hu- whose function in virus replication or packaging remains largely
man populations, and its spread has resulted in a global pandemic uncharacterized (Chen and Zhong, 2020; Finkel et al., 2020).
with more than a million deaths to date (Andersen et al., 2020; Zou As obligate intracellular parasites, viruses require host cell
et al., 2020). We do not fully understand the molecular basis of components to translate and transport their proteins and to
infection and pathogenesis of this virus in human cells. Accord- assemble and secrete viral particles (Maier et al., 2016). The
ingly, there is an urgent need to understand these mechanisms mammalian innate immune system acts to rapidly detect and
to guide the development of therapeutic agents. block viral infection at all stages of the virus life cycle (Chow
SARS-CoV-2 encodes 27 proteins with diverse functional roles et al., 2018; Jensen and Thomsen, 2012; Wilkins and Gale,
in virus replication and packaging (Bar-On et al., 2020; Wang et al., 2010). The primary form of intracellular virus surveillance

Cell 183, 1325–1339, November 25, 2020 ª 2020 The Authors. Published by Elsevier Inc. 1325
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
ll
OPEN ACCESS Article

engages the interferon pathway, which amplifies signals result- that resides in the mRNA entry channel of the initiating 40S ribo-
ing from detection of intracellular viral components to induce a some. This interaction leads to global inhibition of mRNA trans-
systemic type I interferon response upon infection (Stetson lation upon SARS-CoV-2 infection of human cells. Finally, we
and Medzhitov, 2006). Specifically, cells contain various RNA find that NSP8 and NSP9 bind to discrete regions on the 7SL
sensors (such as RIG-I and MDA5) that detect the presence of RNA component of the signal recognition particle (SRP) and
viral RNAs and promote nuclear translocation of the transcription interfere with protein trafficking to the cell membrane upon infec-
factor IRF3, leading to transcription, translation, and secretion of tion. We show that disruption of each of these essential cellular
interferon (e.g., interferon [IFN]-a and IFN-b). Binding of IFN to functions acts to suppress the type I IFN response to viral infec-
cognate cell-surface receptors leads to transcription and trans- tion. Our results uncover a multipronged strategy utilized by
lation of hundreds of antiviral genes. SARS-CoV-2 to antagonize essential cellular processes and
In order to successfully replicate, viruses employ a range of robustly suppress host immune defenses.
strategies to counter host antiviral responses (Beachboard and
Horner, 2016). In addition to their essential roles in the viral life RESULTS
cycle, many viral proteins also antagonize core cellular functions
in human cells to evade host immune responses. For example, Comprehensive Mapping of SARS-CoV-2 Protein
human cytomegalovirus (HCMV) encodes proteins that inhibit Binding to Human RNAs
major histocompatibility complex (MHC) class 1 display on the We cloned all 27 of the known SARS-CoV-2 viral proteins into
cell surface by retaining MHC proteins in the endoplasmic retic- mammalian expression vectors containing an N-terminal Halo-
ulum (Miller et al., 1998), polioviruses encode proteins that Tag (Los et al., 2008; Figure S1A; STAR Methods), expressed
degrade translation initiation factors (eIF4G) to prevent transla- each in HEK293T cells, and exposed them to UV light to cova-
tion of 50 -capped host mRNAs (Kempf and Barton, 2008; Lloyd, lently crosslink proteins to their bound RNAs. We then lysed
2006), and influenza A encodes a protein that modulates mRNA the cells and purified each viral protein using stringent dena-
splicing to degrade the mRNA that encodes RIG-I (Kochs et al., turing conditions to disrupt any non-covalent associations and
2007; Zhang et al., 2018). capture those with a UV-mediated interaction (Figure 1A; STAR
Suppression of the IFN response has recently emerged as a Methods). As positive and negative controls, we purified a known
major clinical determinant of COVID-19 severity (Zhang et al., human RNA binding protein (PTBP1) and a metabolic protein
2020), with almost complete loss of secreted IFN characterizing (GAPDH) (Figures S1A–S1E).
the most severe cases (Hadjadj et al., 2020). The extent to which We successfully purified 26 of the 27 viral proteins (Figure S1A;
SARS-CoV-2 suppresses the IFN response is a key character- full-length S was not soluble when expressed). We found that 10
istic that distinguishes COVID-19 from SARS and Middle East viral proteins (NSP1, NSP4, NSP8, NSP9, NSP12, NSP15,
respiratory syndrome (MERS) (Lokugamage et al., 2020). Several NSP16, ORF3b, N, and E) bind to specific host RNAs (p <
strategies have been proposed for how the related SARS- and 0.001; Figure 1B; Table S1), including 6 structural ncRNAs and
MERS-causing viruses may hijack host cell machinery and 142 mRNAs (Table S1). These include mRNAs involved in protein
evade immune detection, including repression of host mRNA translation (e.g., COPS5, EIF1, and RPS12,), protein transport
transcription in the nucleus (Canton et al., 2018), degradation (ATP6V1G1, SLC25A6, and TOMM20), protein folding (HSPA5,
of host mRNA in the nucleus and cytoplasm (Kamitani et al., HSPA6, and HSPA1B), transcriptional regulation (YY1, ID4, and
2009; Lokugamage et al., 2015), and inhibition of host translation IER5), and immune response (JUN, AEN, and RACK1) (false dis-
(Nakagawa et al., 2018). Nonetheless, the extent to which SARS- covery rate [FDR] < 0.05; Figures 1B and S1F). Importantly, the
CoV-2 uses these or other strategies and how they may be observed interactions are highly specific for each viral protein,
executed at a molecular level remains unclear. and each protein binds to a precise region within each RNA (Fig-
Understanding the interactions between viral proteins and ures 1C and S1F).
components of human cells is essential for elucidating their path- Using these data, we identified several viral proteins that
ogenic mechanisms and for development of effective therapeu- interact with structural ncRNA components of the spliceosome
tic agents. Because SARS-CoV-2 is an RNA virus, and many of (U1 and U2 small nuclear RNA [snRNA]), the ribosome (18S
its encoded proteins are known to bind RNA (Sola et al., 2011), and 28S rRNA), and the SRP (7SL) (Figure 1B). Because these
we reasoned that these viral proteins may interact with specific molecular machines are essential for three essential steps of
human mRNAs (critical intermediates in protein production) or protein production—mRNA splicing, translation, and protein
non-coding RNAs (critical structural components of diverse trafficking—we focused on their interactions with viral proteins
cellular machines) to promote virus propagation. to understand their functions and mechanisms in SARS-CoV-2
Here we comprehensively define the interactions between pathogenesis.
each SARS-CoV-2 protein and human RNA. We show that 10
viral proteins form highly specific interactions with mRNAs or NSP16 Binds to the Pre-mRNA Recognition Domains of
noncoding RNAs (ncRNAs), including those involved in progres- the U1 and U2 snRNAs
sive steps of host cell protein production. We show that NSP16 After transcription in the nucleus, nascent pre-mRNAs are spliced
binds to the mRNA recognition domains of the U1 and U2 RNA to generate mature mRNAs that are translated into protein.
components of the spliceosome and acts to suppress global Splicing is mediated by a complex of ncRNAs and proteins known
mRNA splicing in SARS-CoV-2-infected human cells. We find as the spliceosome. Specifically, the U1 snRNA hybridizes to the 50
that NSP1 binds to a precise region on the 18S ribosomal RNA splice site at the exon-intron junction, and the U2 snRNA

1326 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

A B

Figure 1. Global RNA Binding Maps of SARS-CoV-2 Proteins


(A) Schematic of our approach.
(B) Enrichment heatmap of each SARS-CoV-2 protein (rows) by significantly enriched 100-nt RNA bins (columns; p < 0.001 and enrichment > 3-fold; STAR
Methods). Shared colored bars indicate multiple bins within the same mRNA. For spacing reasons, the 82 mRNAs bound by N protein are displayed separately.
(C) Examples of sequencing reads over specific mRNAs for viral proteins (red) relative to input RNA coverage (gray) are shown. Coding regions (thick lines) and
untranslated regions (thin lines) are shown for each mRNA.
See also Table S1.

hybridizes to the branchpoint site in the intron to initiate splicing of nents, we hypothesized that NSP16 might disrupt splicing of
virtually all human mRNAs (Séraphin et al., 1988). newly transcribed genes (Figure 2F). To test this, we co-ex-
We identified a highly specific interaction between the NSP16 pressed NSP16 in human cells along with a splicing reporter
viral protein and the U1 and U2 snRNAs (Figure 1B). Because U1 derived from IRF7 (an exon-intron-exon minigene) fused to
and U2 are small RNAs (164 and 188 nt, respectively), we noticed GFP (Majumdar et al., 2018). In this system, if the reporter is
strong enrichment of NSP16-associated reads across the entire spliced, then GFP is made; if not, then translation is terminated
length of each. To more precisely define the binding sites, we ex- (via a stop codon present in the first intron), and GFP is not pro-
ploited the well-described tendency of reverse transcriptase to duced (Figure 3A). We observed a more than 3-fold reduction in
preferentially terminate when it encounters a UV-crosslinked pro- GFP levels in the presence of NSP16 compared with a control
tein on RNA (König et al., 2010; Figures 1A and S1D). We deter- human protein (Figures 3B and S3A).
mined that NSP16 binds to the 50 splice site recognition sequence To explore whether NSP16 has a global effect on splicing of
of U1 (Figures 2A, 2B, S2A, and S2B) and the branchpoint recog- endogenous mRNAs, we measured the splicing ratio of each
nition site of U2 (Figures 2C, 2D, S2C, and S2D). These binding gene using nascent RNA sequencing. Specifically, we metaboli-
sites are highly specific to NSP16 relative to all of the other viral cally labeled nascent RNA by feeding cells for 20 min with 5-ethy-
and human proteins (Figures 1B, S2A, and 2C). Consistent with nyl uridine (5EU), purified and sequenced 5EU-labeled RNA, and
its interaction with U1/U2, we observed that NSP16 localizes in quantified the proportion of unspliced fragments spanning the 30
the nucleus upon SARS-CoV-2 infection (Figures 2E, S2E, and splice site of each gene (Figures 3C and S3B). We observed a
S2F) and when expressed in human cells (Figure S2G). global increase in the fraction of unspliced genes in the presence
of NSP16 compared with controls (Figures 3D, S3C, and S3D).
NSP16 Disrupts Global mRNA Splicing upon SARS-CoV- Given that NSP16 is sufficient to suppress global mRNA
2 Infection splicing, we expect that its expression in SARS-CoV-2-infected
Based on the locations of the NSP16 binding sites relative to the cells would result in a global mRNA splicing deficit. To test this,
mRNA recognition domains of the U1/U2 spliceosomal compo- we infected human lung epithelial cells (Calu3) with SARS-CoV-2

Cell 183, 1325–1339, November 25, 2020 1327


ll
OPEN ACCESS Article

A B

C
D

E F

Figure 2. NSP16 Binds to U1 and U2 at Their mRNA Recognition Sites


(A) NSP16 enrichment of reverse transcription stop positions across each nucleotide of U1 (red) compared with a control protein (GAPDH, black). The red box
(below the x axis) represents most enriched nucleotide positions (U1, 9–13 nt). The gray-shaded box (overlay) outlines the position of the splice site recognition
sequence.
(B) Left: structure of the pre-catalytic human spliceosome (PDB: 6QX9; Charenton et al., 2019), highlighting the location of NSP16 binding site (red spheres)
relative to U1 (yellow ribbon) and mRNA (purple ribbon). Right: schematic of the structure.
(C) Enrichment across each nucleotide of U2 for NSP16 (red) and GAPDH (black). The red box demarcates most enriched nucleotide positions (U2, 27–34 nt). The
gray-shaded box outlines the location of the branchpoint recognition sequence.
(D) Structure of the pre-catalytic human spliceosome (PDB: 6QX9; Charenton et al., 2019) displaying the NSP16 binding site (red spheres), U2 (orange), and
mRNA (purple).
(E) Mock-infected (top) or SARS-CoV-2 infected (bottom) Vero E6 cells immunostained with a polyclonal antibody to NSP16 (left) or NSP1 (right). Imaris 3D
reconstruction of the DAPI (nucleus) and NSP16 or NSP1 signal are shown for each protein. The signal contained within the 3D nuclear volume (blue) is shown in
yellow and the cytoplasmic signal in purple. Scale bars, 3 mm.
(F) Model: NSP16 binding to U1/U2 can affect mRNA recognition during splicing.

and measured the splicing levels of newly transcribed mRNAs to act as an enzyme that deposits 20 -O-methyl modifications on
compared with a mock-infected control. As expected, we viral RNAs (Decroly et al., 2011), our results demonstrate that it
observed a global increase in the fraction of unspliced tran- also acts as a host virulence factor. Global disruption of mRNA
scripts upon SARS-CoV-2 infection, with 90% of measured splicing may act to decrease host protein and mRNA levels by trig-
genes showing increased intron retention (Figures 3E and S3E). gering nonsense-mediated decay of improperly spliced mRNAs
These results indicate that NSP16 binds to the splice site and (Kurosaki et al., 2019). Consistent with this, we observed a strong
branchpoint sites of U1/U2 to suppress global mRNA splicing in global decrease in steady-state mRNA levels (relative to ncRNA
SARS-CoV-2-infected cells (Figure 3F). Although NSP16 is known levels) upon SARS-CoV-2 infection (Figure S3F).

1328 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

A B

D E

F G H

Figure 3. NSP16 Suppresses Host mRNA Splicing


(A) Schematic of fluorescence reporter used to assay mRNA splicing.
(B) GFP density plot of HEK293T cells expressing the GFP splicing reporter and either GAPDH (gray) or NSP16 (red).
(C) Schematic of the nascent RNA purification method.
(D) The percentage of unspliced difference for each gene between HEK293T cells transfected with GAPDH (gray) or NSP16 (red). The plot represents the merge of
four independent biological replicates; replicates are plotted in Figure S4C.
(E) Violin plot for SARS-CoV-2 infected human lung epithelial cells (MOI = 0.01, 48 h) compared with mock infection. Plots are merges of two biological replicates;
replicates are plotted in Figure S4E.
(F) Model. NSP16 binding to U1 and U2 can reduce overall mRNA and protein levels.
(G) Expression of an IFN-stimulated gene (ISG) reporter upon transfection with GAPDH (gray) or NSP16 (red) after stimulation with IFN-b. Three independent
biological replicates; **p < 0.01.
(H) Example of nascent RNA sequencing at the intron of ISG15 (intron, line; exon, box) upon SARS-CoV-2 (red) or mock (gray) infection.

Inhibition of mRNA Splicing Suppresses the Host IFN an antiviral response gene). This IFN-stimulated gene (ISG) re-
Response to Viral Infection porter line can be stimulated using IFN-b and assayed for re-
Because many of the key genes stimulated by IFN are spliced, porter induction. We observed strong repression of this IFN-
we reasoned that mRNA splicing would be critical for a robust responsive gene upon expression of NSP16 (Figure 3G) and
IFN response. To test this, we utilized a reporter line engineered upon addition of a small molecule that interferes with spliceoso-
to express alkaline phosphatase upon IFN signaling (mimicking mal assembly (Figure S3G). These results demonstrate that one

Cell 183, 1325–1339, November 25, 2020 1329


ll
OPEN ACCESS Article

A B

C D

E F G

H I J

Figure 4. NSP1 Binds to 18S Near the mRNA Entry Channel to Suppress Translation
(A) NSP1 enrichment across each nucleotide of 18S. The cyan box indicates the most enriched nucleotides of NSP1 binding (18S, 607–644 nt).
(B) The location of NSP1 binding (cyan spheres) relative to the known structure of 40S (gray) and mRNA (purple ribbon). Right: schematic illustrating structure
(Ameismeier et al., 2018) and how NSP1 binding would block mRNA entry.
(C) Images of HEK293T cells co-expressing the GFP reporter and GAPDH (top) or NSP1 (bottom).
(D) Flow cytometry quantification (mean intensity) of GFP in the presence of GAPDH, NSP8/9, M, or NSP1 proteins. Three independent biological replicates per
condition.
(E) Puromycin incorporation (top) or total actin levels (bottom) measured in Calu3 cells infected with SARS-CoV-2 (MOI = 0.01, 48 h) or a mock-infected control
(left 2 lanes).
(F) The ratio of puromycin signal over total actin signal is plotted for each individual replicate.
(G) Read enrichment on 18S for an independent replicate of NSP1 wild type, NSP1 R124A/K125A mutant, and NSP1 K164A/H165A (DRC) mutant.
(H) Flow cytometry analysis of HEK293T cells transfected with GFP and NSP1DRC mutant (gray), wild-type NSP1, or NSP1 R124A/K125A (cyan).

(legend continued on next page)

1330 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

outcome of NSP16-mediated inhibition of mRNA splicing is to (Figures 4C and 4D). In contrast, we did not observe this inhibition
reduce the host cells’ innate immune response to virus recogni- when we expressed other SARS-CoV-2 proteins (NSP8, NSP9, or
tion. Consistent with such a role, we observed an increase in M) or human proteins (GAPDH) (Figure 4D).
intron retention in multiple IFN-responsive genes (such as To determine whether NSP1 leads to translational inhibition of
ISG15 and RIG-I) upon SARS-CoV-2 infection (Figures 3H, endogenous proteins in human cells, we used a technique called
S3H, and S3I). surface sensing of translation (SUnSET) to measure global pro-
tein production levels (Schmidt et al., 2009). In this assay, trans-
NSP1 Binds to 18S Ribosomal RNA in the mRNA Entry lational activity is measured by the level of puromycin incorpora-
Channel of the 40S Subunit tion into elongating polypeptides (Figure S4E). We observed a
When exported to the cytoplasm, spliced mRNA is translated strong reduction in the level of global puromycin integration in
into protein on the ribosome. Initiation of translation begins cells expressing NSP1 compared with cells expressing GFP (Fig-
with recognition of the 50 cap by the small 40S subunit (which ures S4F and S4G).
scans the mRNA to find the first start codon). We observed Because NSP1 expression is sufficient to suppress global
that NSP1 binds exclusively to the 18S ribosomal RNA (Figures mRNA translation in human cells, we hypothesized that SARS-
1B and S4A)—the structural RNA component of the 40S ribo- CoV-2 infection would also suppress global translation. To test
somal subunit. this, we infected a human lung epithelial (Calu3) or monkey kid-
Several roles of NSP1 have been reported in SARS-CoV and ney (Vero) cell line with SARS-CoV-2 and measured nascent pro-
MERS-CoV, including roles in viral replication, translational inhi- tein synthesis levels using SUnSET. We observed a strong
bition, transcriptional inhibition, mRNA degradation, and cell cy- reduction of global puromycin integration upon SARS-CoV-2
cle arrest (Brockway and Denison, 2005; Kamitani et al., 2009; infection in both cell types (Figures 4E, 4F, S4H, and S4I).
Lokugamage et al., 2015; Narayanan et al., 2015). One of the re- To explore whether NSP1 binding to 18S rRNA is critical for
ported roles of NSP1 in SARS-CoV is that it can associate with translational repression, we generated a mutant NSP1 in which
the 40S ribosome to inhibit host mRNA translation (Kamitani two positively charged amino acids (K164 and H165) in the C-ter-
et al., 2009; Tanaka et al., 2012), but it remains unknown whether minal domain were replaced with alanine residues (Figure S4C;
this association is due to interaction with the ribosomal RNA, Narayanan et al., 2008). We observed complete loss of in vivo con-
protein components of the ribosome, or other auxiliary ribosomal tacts with 18S (Figure 4G); because this mutant disrupts ribosome
factors. Accordingly, the mechanisms by which NSP1 acts to contact, we refer to it as NSP1DRC. We co-expressed GFP and
suppress protein production remain elusive. NSP1DRC in HEK293T cells and found that the mutant fails to
We mapped the location of NSP1 binding to a 37-nt region inhibit translation (Figures 4H and S4J). In contrast, mutations to
corresponding to helix 18 (Figure 4A), adjacent to the mRNA en- the positively charged amino acids at positions 124/125 do not
try channel (Simonetti et al., 2020; Figure 4B). The interaction affect 18S binding (Figure 4G) or the ability to inhibit translation
would position NSP1 to disrupt 40S mRNA scanning and prevent (Figure 4H).
translation initiation (Figure 4B) and disrupt tRNA recruitment to These results demonstrate that NSP1 binds in the mRNA entry
the 80S ribosome and block protein production (Figure S4B). channel of the ribosome and that this interaction is required for
Interestingly, the NSP1 binding site includes the highly translational inhibition of host mRNAs upon SARS-CoV-2
conserved G626 nucleotide, which monitors the minor groove infection.
of the codon-anticodon helix for tRNA binding fidelity (Ogle
et al., 2001). We noticed that the C-terminal region of NSP1 NSP1-Mediated Translational Inhibition Suppresses the
has structural regions similar to SERBP1 (Brown et al., 2018) Host IFN Response
and Stm1 (Ben-Shem et al., 2011), two known ribosome inhibi- We explored whether NSP1 binding to 18S rRNA suppresses the
tors that bind in the mRNA entry channel to preclude mRNA ac- ability of cells to respond to IFN-b stimulation upon viral infec-
cess (Figure S4C). Consistent with this, a recent cryo-EM struc- tion. We transfected ISG reporter cells with NSP1, stimulated
ture confirms that NSP1 binds to these same nucleotides of 18S with IFN-b, and observed robust repression of the IFN-respon-
within the mRNA entry channel (Thoms et al., 2020). sive gene (>6-fold; Figure 4I). To confirm that this NSP1-medi-
ated repression occurs in human cells upon activation of dou-
NSP1 Suppresses Global Translation of Host mRNAs ble-stranded RNA (dsRNA)-sensing pathways typically
upon SARS-CoV-2 Infection triggered by viral infection, we treated a human lung epithelial
Given the location of NSP1 binding on the 40S ribosome, we hy- cell line (A549) with poly(I:C), a molecule that is structurally
pothesized that it could suppress global initiation of mRNA trans- similar to dsRNA and known to induce an antiviral innate immune
lation. To test this, we performed in vitro translation assays of a response (Alexopoulou et al., 2001; Kato et al., 2006) (Fig-
GFP reporter in HeLa cell lysates and found that addition of ure S4K). We observed marked downregulation of IFN-b protein
NSP1 led to potent inhibition of translation (Figure S4D). We and endogenous IFN-b-responsive mRNAs in the presence of
observed a similar NSP1-mediated translational repression when NSP1 but not in the presence of NSP1DRC (Figures S4L and
we co-expressed NSP1 and a GFP reporter gene in HEK293T cells S4M). These results demonstrate that NSP1, through its

(I) Quantification of the IFN-b response in the presence of GAPDH (gray) or NSP1 (cyan).
(J) Schematic of how NSP1 acts to suppress mRNA translation.
Error bars represent standard deviation across biological replicates, and dots represent individual values for each replicate; *p < 0.05 and **p < 0.01.

Cell 183, 1325–1339, November 25, 2020 1331


ll
OPEN ACCESS Article

A B C

D E F

Figure 5. The 50 Viral Leader Protects mRNA from NSP1-Mediated Translational Inhibition
(A) Images of cells co-transfected with NSP1 and mCherry alone ( leader, top) or mCherry fused to the SARS-CoV-2 leader (+ leader, bottom).
(B) GFP (green) or mCherry (red) levels when fused to the viral leader (+ leader, right) or lacking the viral leader ( leader, left).
(C) GFP reporter with no leader (left), full leader (center), or stem loop 1 (SL1) upon NSP1 expression.
(D) Calu3 cells expressing SL1 fused to GFP. Cells were mock or SARS-CoV-2 infected (MOI = 0.1), and GFP expression was measured 24 h after infection by flow
cytometry.
(E) GFP reporter containing SL1 (left), a swap of SL2 and SL1 (SL2-SL1), insertion of 5 nt between the 50 end and SL1 (+5 nt-SL1), or no leader. GFP protein level
was measured for each condition upon expression of NSP1.
(F) Proposed model of how NSP1 binding to the viral leader can allosterically modulate NSP1 structure to protect mRNAs in cis.
Error bars represent standard deviation across biological replicates, and dots represent individual replicate values; *p < 0.05 and **p < 0.01.

interaction with 18S rRNA, suppresses the innate immune ing, (2) it could directly recruit free ribosomes, or (3) NSP1 could
response to virus recognition (Figure 4J). bind to the leader independent of its ribosome interaction to alloste-
rically modulate the NSP1-ribosome interaction. We reasoned that
The Viral 50 Leader Protects mRNA from NSP1-Mediated if the leader competes for NSP1 binding or directly recruits free ri-
Translational Inhibition bosomes, then the presence of SL1 should be sufficient for protec-
Because NSP1 blocking the mRNA entry channel would affect tion, regardless of its precise position in the 50 UTR. In contrast, if
host and viral mRNA translation, we explored how translation the leader allosterically modulates ribosome binding, then the
of viral mRNAs is protected from NSP1-mediated translational spacing between the 50 cap (which is bound to NSP1-40S) and
inhibition. Many viruses contain 50 untranslated regions that SL1 would be critical for protection. To distinguish between these
regulate viral gene expression and translation (Gaglia et al., models, we swapped the location of SL1 and SL2 in the 50 leader
2012); all SARS-CoV-2-encoded subgenomic RNAs contain a or inserted 5 nt between the 50 cap and SL1 (Figure S5C) and found
common 50 leader sequence that is added during negative- that both mutants ablate protection (Figures 5E and S5D).
strand synthesis (Kim et al., 2020b). We explored whether the These results indicate that an mRNA requires the 50 leader to
leader sequence protects viral mRNAs from translational inhibi- be precisely positioned relative to the NSP1-bound 40S ribo-
tion by fusing the viral leader sequence to the 50 end of GFP or some to enable translational initiation (Figure 5F). Although
mCherry reporter genes (Figure S5A). We found that NSP1 fails many aspects of this allosteric model remain to be explored, it
to suppress translation of these leader-containing mRNAs (Fig- would explain how leader-mediated protection can occur on
ures 5A, 5B, and S5B). We dissected the leader sequence and an mRNA only when present in cis. Moreover, this model sug-
found that the first stem loop (SL1) is sufficient to prevent trans- gests that NSP1 might also act to further increase viral mRNA
lational suppression upon NSP1 expression (Figure 5C) or translation by actively recruiting the ribosome to its own mRNAs.
SARS-CoV-2 infection (Figure 5D). Consistent with this, we observed a consistent, 20% increase
We considered three models for how the leader could protect in translation of leader-containing reporter levels upon viral infec-
viral mRNAs: (1) it could compete with the ribosome for NSP1 bind- tion (Figure 5D) or expression of NSP1 (Figure S5E).

1332 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

A B

Figure 6. NSP8 and NSP9 Bind to 7SL RNA of the SRP


(A) Enrichment of reverse transcription stop positions across each nucleotide of 7SL is shown for NSP8 (blue) and NSP9 (red). Red (7SL, 142–143 nt; 7SL, 149–
151 nt) and blue (7SL, 193–194 nt) boxes demarcate the most enriched nucleotide positions.
(B) The locations of the NSP8 (blue spheres) and NSP9 (red spheres) binding sites on the S domain of 7SL (yellow ribbon) structure relative to SRP54 and SRP19
(gray) (PDB: 1MFQ; Kuglstatter et al., 2002). Right: schematic of the structure and model of how NSP8/9 binding to 7SL could affect SRP protein binding.
(C) Read enrichment across each nucleotide of 28S for NSP8 (blue). The black box indicates the location of the ES27 expansion sequence (28S, 2,889–3,551 nt).
The blue box indicates the most enriched nucleotide position on 28S rRNA (28S, 3,017–3,529 nt).
(D) The locations of the NSP8 (blue) and NSP9 (red) binding sites relative to the structure of the SRP-ribosome complex (PDB: 3JAJ; Voorhees and Hegde, 2015)
superimposed on the structure of the ES27 region of 28S (Ebp1-ribosome complex; PDB: 6SXO; Wild et al., 2020). The observed NSP8 binding site in the ES27
region of 28S (gray) is demarcated in blue, and the NSP8 (blue) and NSP9 (red) binding sites on 7SL (yellow) are highlighted. Right: schematic illustrating the
interaction between the ribosome and SRP.

NSP8 and NSP9 Bind to the 7SL RNA Component of SRP for signal peptide recognition, SRP-receptor binding, and ribo-
Upon engaging the start codon in an mRNA, the 60S subunit of some translocation; Akopian et al., 2013; Holtkamp et al.,
the ribosome is recruited to form the 80S ribosome, which trans- 2012; Figure 6B). NSP9 binds to 7SL in the region that is bound
lates mRNA. SRP is a universally conserved complex that binds by the SRP19 protein (Figure 6B), which is required for proper
to the 80S ribosome and acts to co-translationally scan the folding and assembly of the SRP (including proper loading of
nascent peptide to identify hydrophobic signal peptides present SRP54; Akopian et al., 2013).
in integral membrane proteins and proteins secreted from the Because SRP scans nascent peptides co-translationally, we
plasma membrane (Akopian et al., 2013). When these are identi- were intrigued to find that NSP8 also forms a highly specific
fied, SRP triggers ribosome translocation to the endoplasmic re- interaction with 28S rRNA (the structural component of the 60S
ticulum (ER) to ensure proper folding and trafficking of these pro- subunit) (Figures 6C and S6B). The binding site on 28S rRNA cor-
teins to the cell membrane (Akopian et al., 2013). responds to the largest human-specific expansion segment in
We identified two viral proteins, NSP8 and NSP9, that bind at the ribosome, referred to as ES27 (Parker et al., 2018). ES27 is
distinct and highly specific regions in the S domain of the 7SL highly dynamic and, thus, has not been resolved in most ribo-
RNA scaffold of SRP (Figures 6A and S6A). NSP8 interacts some structures (Zhang et al., 2014). However, when engaged
with 7SL in the region bound by SRP54 (the protein responsible by specific factors, ES27 can become ordered and has been

Cell 183, 1325–1339, November 25, 2020 1333


ll
OPEN ACCESS Article

shown recently to be capable of interacting with the ribosome shift in membrane protein levels only occurred in S+ cells (Fig-
exit tunnel adjacent to the 60S binding site of SRP (Figures 6D ure 7D), whereas the S- population resembled the mock-infected
and S6C; Wild et al., 2020). samples (Figure 7C). We observed a strong relationship between
These observations suggest that NSP8 and NSP9 bind to the the level of S protein, likely reflecting the amount of viral replica-
co-translational SRP complex. Consistent with this, we find that tion in each cell, and the level of membrane protein suppression
NSP8 and NSP9 localize broadly throughout the cytoplasm (Figure 7C). We observed this membrane protein-specific
when expressed in human cells (Figure S6D) or upon SARS- decrease upon infection of human lung epithelial cells (Calu3;
CoV-2 infection (Figures S6E and S6F). Figure S7D) and monkey kidney cells (Vero; Figures 7C and 7D).
These results demonstrate that NSP8 and NSP9 bind to 7SL to
NSP8 and NSP9 Suppress Protein Integration into the disrupt SRP function and suppress membrane protein trafficking
Cell Membrane in SARS-CoV-2-infected cells. Although NSP8 and NSP9 are
Because NSP8 and NSP9 binding on 7SL are positioned to thought to be components of the virus replication machinery
disrupt SRP function, we hypothesized that they may alter trans- (Sutton et al., 2004), our results indicate that they play an addi-
location of secreted and integral membrane proteins tional role as host virulence factors. Because viral membrane
(Figure S7A). proteins also require trafficking to the ER, viral disruption of
To test this, we expressed an SRP-dependent membrane pro- SRP might also negatively impact virus propagation, unless viral
tein (nerve growth factor receptor [NGFR]; Izon et al., 2001) fused proteins are trafficked in an SRP-independent manner (Fig-
via an internal ribosome entry site (IRES) to a non-membrane ure S7E) or if NSP8/9 selectively affects host (but not viral) pro-
GFP (Figure S7F). In this system, when a perturbation specifically tein trafficking.
affects membrane protein levels, we expect to see a decrease in
the ratio of membrane to non-membrane protein levels. To Viral Disruption of Protein Trafficking Suppresses the
ensure that the NGFR reporter accurately reports SRP function, IFN Response
we treated HEK293T cells with small interfering RNAs (siRNAs) Next we explored how disruption of SRP might be advantageous
against SRP54 or SRP19 and found that both lead to a dramatic for virus propagation. Because secretion of IFN and other cyto-
reduction of the NGFR membrane protein relative to the non- kines is dependent on the SRP complex for secretion (Fig-
membrane GFP protein (Figure S7B). Similarly, we found that ure S7F), a central component of the IFN response is dependent
expression of NSP8 and NSP9 (alone or together) led to a striking on SRP. Accordingly, we hypothesized that NSP8/9-mediated
reduction in expression of NGFR relative to GFP (Figure 7A). viral suppression of SRP would act to suppress the IFN response
Expression of control proteins did not specifically affect NGFR upon infection. To test this, we co-expressed NSP8 and NSP9
levels (Figures 7A and S7B). and observed a significant reduction in the IFN response relative
To determine whether there is a global effect on membrane to a control protein (Figure S7G).
protein levels, we utilized the SUnSET method to measure puro- These results suggest that SARS-CoV-2-mediated suppres-
mycin levels in membrane proteins using flow cytometry (STAR sion of SRP-dependent protein secretion enables suppression
Methods). We confirmed that disruption of SRP leads to a global of host immune defenses (Figure 7E). Interestingly, many pro-
reduction in puromycin levels in the cell membrane (Figure S7C). teins involved in anti-viral immunity, including most cytokines
We observed a comparable global reduction of puromycin- and MHC class I, are membrane anchored or secreted and are
labeled membrane proteins upon expression of NSP8 or NSP9 known to use the SRP pathway for transport (Vermeire et al.,
individually or together but not with control proteins (Figures 2014; Figure S7F), suggesting that there may be other effects
7B and S7C). of SRP pathway inhibition on SARS-CoV-2 pathogenesis.

SARS-CoV-2 Infection Suppresses Protein Integration DISCUSSION


into the Cell Membrane
Because NSP8 and NSP9 are each sufficient to suppress protein We identified several pathogenic functions of SARS-CoV-2 in hu-
integration into the cell membrane, we anticipated that SARS- man cells, including global inhibition of host mRNA splicing, pro-
CoV-2 infection would lead to similar suppression. However, tein translation, and membrane protein trafficking, and described
determining whether SARS-CoV-2 infection specifically affects the molecular mechanisms by which the virus acts to disrupt these
membrane protein expression is confounded by the fact that essential cell processes. Interestingly, all of the viral proteins
NSP1 inhibits translation of membrane and non-membrane pro- involved (NSP1, NSP8, NSP9, and NSP16) are produced in the
teins upon infection. first stage of the virus life cycle prior to generation of dsRNA prod-
To address this, we co-expressed a membrane protein re- ucts during virus genome replication. Because dsRNA is detected
porter (NGFR) containing the 50 viral leader along with a non- by host immune sensors and triggers the type I IFN response,
membrane GFP reporter containing the viral leader. Upon viral disruption of these cellular processes allows the virus to replicate
infection, we observed a strong reduction in membrane protein its genome while minimizing the host innate immune response.
levels (Figure 7C) but no reduction in non-membrane GFP levels Disruption of these three non-overlapping steps of protein pro-
(Figure 5D). To ensure that these effects are specific to SARS- duction may represent a multi-pronged mechanism that synergis-
CoV-2-infected cells, we separated individual cells in the in- tically acts to suppress the host antiviral response (Figure 7F). Spe-
fected population into those expressing the viral S protein (S+) cifically, the IFN response is usually boosted more than 1,000-fold
and those not expressing the protein (S). We found that the upon virus detection (through amplification and feedback;

1334 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

A B E

C D

Figure 7. NSP8 and NSP9 Inhibit Membrane and Secretory Protein Trafficking
(A) Quantification of HEK293T cells transfected with plasmids co-expressing GFP-tagged NSPs and the NGFR membrane protein. Plotted is the ratio of NGFR to
GFP levels for each condition.
(B) The ratio of puromycin-containing proteins at the cell membrane normalized to GFP expression for each condition.
(C) Quantification of two mRNA reporters containing SL1 fused to GFP (leader-GFP) or NGFR (leader-NGFR) in Vero cells infected with SARS-CoV-2 or mock
infected for 24 h (MOI, 0.1). Plotted is the ratio of leader-NGFR to leader-GFP, binned by increasing amounts of S protein.
(D) Density plot for leader-NGFR to leader-GFP ratios in virally infected Vero cells or mock-treated controls. Replicate conditions were merged for display.
(E) Model of how NSP8/9 act to suppress SRP-dependent protein trafficking upon viral infection.
(F) A model of how SARS-CoV-2 suppresses host immune responses through multi-pronged inhibition of core cellular functions. Cellular mechanisms are shown
in gray and viral mechanisms in red.
Error bars represent standard deviation across independent biological replicates, and dots represent individual values for each replicate; *p < 0.05 and **p < 0.01.

Cell 183, 1325–1339, November 25, 2020 1335


ll
OPEN ACCESS Article

Figure S4K), but each individual mechanism affects IFN levels 5- Accordingly, it remains possible that our maps may not fully cap-
to 10-fold. Accordingly, if each independent mechanism moder- ture all of the interactions that occur when human cells are in-
ately affects IFN levels, the three together may be able to achieve fected, such as interactions that occur with virus-induced
dramatic suppression of IFN (103 = 1,000-fold). This multi-pronged RNAs, in specific viral compartments, or that require multiple
mechanism may explain the molecular basis of the potent sup- viral proteins. (2) Although we characterized the functional and
pression of IFN observed in patients with severe COVID-19. mechanistic roles of several viral proteins and structural
IFN is emerging not only as a determinant of disease severity ncRNAs, we did not explore the roles of viral protein interactions
but also as a potential treatment option (Zhou et al., 2020). Our with mRNAs. (3) How the virus disrupts fundamental cellular pro-
work identifies several therapeutic opportunities for boosting cesses while maintaining its own production is still largely unde-
IFN levels upon SARS-CoV-2 infection. For example, disrupting fined. Although we showed that the 50 leader is sufficient to
the interaction between NSP1 and 18S rRNA could allow cells to relieve translational inhibition by NSP1, we still do not fully under-
detect and respond to viral infection. Because many small-mole- stand how this protection occurs and, specifically, how NSP1
cule drugs target ribosomal RNAs (Liaud et al., 2019), it may be might interact with the viral leader or allosterically modulate ribo-
possible to develop drugs to block NSP1-18S and other interac- some binding. Similarly, viral membrane proteins are dependent
tions. Additionally, disrupting the 50 viral leader may be a potent on trafficking to the ER, and how NSP8/9 might selectively affect
antivirus strategy because it is critical for translation of all viral ER translocation of host but not viral proteins remains to be
proteins. Because SL1 is a structured RNA, it may be possible explored. (4) Although we showed that viral disruption of these
to design small molecules that specifically bind this structure essential cellular functions can suppress IFN, other roles of
to suppress viral protein production (Hermann, 2016). host cell shutdown in viral pathogenesis and in suppressing
Viral suppression of these cellular functions is not exclusive to other aspects of antiviral immunity, including possible roles in
the IFN response and also affects other spliced, translated, adaptive immune responses, have not been explored.
secreted, and membrane proteins. Many proteins involved in
anti-viral immunity are spliced and/or membrane anchored or STAR+METHODS
secreted; for example, MHC class I, critical for antigen presenta-
tion to CD8 T cells at the cell surface of infected cells (Hansen and Detailed methods are provided in the online version of this paper
Bouvier, 2009). By antagonizing membrane trafficking, SARS- and include the following:
CoV-2 may prevent viral antigens from being presented on MHC
and allow infected cells to escape T cell recognition and clear- d KEY RESOURCES TABLE
ance. In this way, interference with these essential cellular pro- d RESOURCE AVAILABILITY
cesses might further aid SARS-CoV-2 to evade the host immune B Lead Contact
response. B Materials Availability
More generally, we expect that insights gained from the B Data and Code Availability
SARS-CoV-2 protein-RNA binding maps will be critical for d EXPERIMENTAL MODEL AND SUBJECT DETAILS
exploring additional viral mechanisms. Specifically, we identified B Cell lines used in this study
many other interactions, including highly specific interactions B Cell culture conditions
with mRNAs. For example, NSP12 binds to the JUN mRNA (Fig- B SARS-CoV-2 Viral Infection strains and conditions
ure S1E), which encodes the critical immune transcription factor d METHOD DETAILS
c-Jun, which is activated in response to multiple cytokines and B Generation of RNA binding maps
immune signaling pathways (Weston and Davis, 2007). We also B Antibody Generation
identified an interaction between NSP9 and the start codon of B Microscopy imaging
the mRNA that encodes COPS5 (Figure 1C), the enzymatic sub- B Structure modeling
unit of the COP9 signalosome complex, which regulates protein B Recombinant NSP1 production
homeostasis (Cope and Deshaies, 2003), suggesting that it B In vitro translation assays
might disrupt its translation. Interestingly, COPS5 (also known B In vivo translation assays
as JAB1) is known to bind and stabilize c-Jun protein (Claret B SUnSET assay
et al., 1996), and several viruses are known to disrupt this protein B SUnSET in SARS-CoV-2 infected cells
(Lungu et al., 2008; Oh et al., 2006; Tanaka et al., 2006). Although B Membrane protein reporter experiments
it remains unknown what, if any, role these interactions have in B siRNA experiments for SRP19 and SRP54
virus-infected cells, the specificity suggests that they may pro- B Leader-NGFR measurements
vide a selective advantage for virus propagation. B Membrane SUnSET assay
Our results demonstrate that global mapping of RNA binding B Splicing assessment experiments
by viral proteins could enable rapid characterization of mecha- B Interferon stimulation experiments
nisms of emerging pathogenic RNA viruses. B 50 viral leader experiments
B Alignments and phylogeny reconstructions
Limitations of Study d QUANTIFICATION AND STATISTICAL ANALYSIS
Several limitations of our current study will need to be explored in B Analysis of SARS-CoV-2 viral protein binding to RNA
future work. (1) Our mapping experiments were performed with B Splicing analysis of 5EU data
uninfected human cells expressing tagged viral proteins. B Analysis of total RNA in SARS-CoV-2 infected samples

1336 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

SUPPLEMENTAL INFORMATION Andersen, K.G., Rambaut, A., Lipkin, W.I., Holmes, E.C., and Garry, R.F.
(2020). The proximal origin of SARS-CoV-2. Nat. Med. 26, 450–452.
Supplemental Information can be found online at https://doi.org/10.1016/j. Bar-On, Y.M., Flamholz, A., Phillips, R., and Milo, R. (2020). SARS-CoV-2
cell.2020.10.004. (COVID-19) by the numbers. eLife 9, e57309.
Beachboard, D.C., and Horner, S.M. (2016). Innate immune evasion strategies
ACKNOWLEDGMENTS of DNA and RNA viruses. Curr. Opin. Microbiol. 32, 113–119.
Ben-Shem, A., De Loubresse, N.G., Melnikov, S., Jenner, L., Yusupova, G.,
We thank Fritz Roth for clones; Marko Jovanovic, Bil Clemons, Shu-ou Shan, and Yusupov, M. (2011). The structure of the eukaryotic ribosome at 3.0 Å res-
and Jamie Wangen for guidance and discussions; Joanna Jachowicz for ana- olution. Science 334, 1524–1529.
lyses; Shawna Hiley for editing; Inna-Marie Strazhnik for figures; and Scott
Berman, H.M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T.N., Weissig, H.,
Tighe, Pheobe Laaguiby, Roxana del Rio-Guerra, Joyce Oetjen, Philip Eisen-
Shindyalov, I.N., and Bourne, P.E. (2000). The Protein Data Bank. Nucleic
hauer, Arvind H. Patel, and Arthur Wickenhagen for technical help. This work
Acids Res. 28, 235–242.
was supported by NIH F30HL136080 (to A.K.B.); P20GM125498 and
P30GM118228-04 (to E.A.B.); NIH GM7616-40 and F30CA247447 (to P.B.); Bolger, A.M., Lohse, M., and Usadel, B. (2014). Trimmomatic: a flexible
UM1HL120877 and U54GM115516 (to D.M.); R41AI132047 and trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120.
U01AI1141997 (to J.W.B.); U01 DA040612 and U01 HL130007 (to M.G.); the Brockway, S.M., and Denison, M.R. (2005). Mutagenesis of the murine hepa-
USC MD/PhD Program (to A.K.B.); UCLA-Caltech MSTP (to P.B.); American titis virus nsp1-coding region identifies residues important for protein process-
Cancer Society PF-17-240-01 (to N.O.); Wellcome Trust 201366/Z/16/Z (to ing, viral RNA synthesis, and viral replication. Virology 340, 209–223.
S.J.R.); the Office of the Vice President for Research at UVM (to J.W.B.); Her- Brown, A., Baird, M.R., Yip, M.C.J., Murray, J., and Shao, S. (2018). Structures
itage Medical Research Institute (to M.G. and R.V.); and NYSCF, CZI, and Cal- of translationally inactive mammalian ribosomes. eLife 7, e40486.
tech (to M.G.). M.G. is a NYSCF-Robertson Investigator.
Canton, J., Fehr, A.R., Fernandez-Delgado, R., Gutierrez-Alvarez, F.J., San-
chez-Aparicio, M.T., Garcı́a-Sastre, A., Perlman, S., Enjuanes, L., and Sola,
AUTHOR CONTRIBUTION I. (2018). MERS-CoV 4b protein interferes with the NF-kB-dependent innate
immune response during infection. PLoS Pathog. 14, e1006838.
A.K.B. conceived the project and performed protein and 5EU purification,
Charenton, C., Wilkinson, M.E., and Nagai, K. (2019). Mechanism of 50 splice
sequencing, NSP1 and 50 leader flow cytometry, and analyses. M.R.B. devel-
site transfer for human spliceosome activation. Science 364, 362–367.
oped and optimized viral protein purification methods, optimized and per-
formed translation assays and splicing assays, and led development and Chen, L., and Zhong, L. (2020). Genomics functional analysis and drug
execution of biochemical assays and functional experiments. E.A.B. led virus screening of SARS-CoV-2. Genes Dis. Published online April 14, 2020.
work, designed and performed experiments, and provided guidance and sup- https://doi.org/10.1016/j.gendis.2020.04.002.
port regarding analyses, ideas, and discussions of the paper. D.D.H. per- Chow, K.T., Gale, M., Jr., and Loo, Y.-M. (2018). RIG-I and Other RNA Sensors
formed NSP1 translational assays and imaging experiments. L.M.C. per- in Antiviral Immunity. Annu. Rev. Immunol. 36, 667–694.
formed 50 leader experiments. A.C. performed membrane reporter and Claret, F.X., Hibi, M., Dhut, S., Toda, T., and Karin, M. (1996). A new group of
puromycin experiments. P.B. developed 5EU selection and splicing analyses. conserved coactivators that increase the specificity of AP-1 transcription fac-
N.O., S.A.Q., A.E.L., R.V., and M.G. performed analyses. D.G.S., D.G., G.D.L., tors. Nature 383, 453–457.
and S.J.R. generated viral antibodies and performed viral infection. C.L. per- Cope, G.A., and Deshaies, R.J. (2003). COP9 signalosome: a multifunctional
formed imaging and analysis. J.T., Z.D.M., and M.M.S. performed experi- regulator of SCF and other cullin-based ubiquitin ligases. Cell 114, 663–671.
ments. R.M.V. provided guidance regarding ribosome and SRP structure
da Silva, S.J.R., da Silva, C.T.A., Mendes, R.P.G., and Pena, L. (2020). Role of
and function. J.W.B. provided guidance and reagents for virus work. D.M.
Nonstructural Proteins in the Pathogenesis of SARS-CoV-2. J. Med. Virol.
and M.G. led the project, supervised experiments and analyses, and wrote
Published online April 9, 2020. https://doi.org/10.1002/jmv.25858.
the paper.
Decroly, E., Debarnot, C., Ferron, F., Bouvet, M., Coutard, B., Imbert, I., Gluais,
L., Papageorgiou, N., Sharff, A., Bricogne, G., et al. (2011). Crystal structure
DECLARATION OF INTERESTS
and functional analysis of the SARS-coronavirus RNA cap 20 -O-methyltrans-
ferase nsp10/nsp16 complex. PLoS Pathog. 7, e1002059.
The authors declare no competing interests.
Delano, W. (2002). PyMOL molecular graphics system. https://pymol.org/2/.
Received: June 18, 2020 Dobin, Alexander, Davis, C.A., Schlesinger, F., Drenkow, J., Zaleski, C., Jha,
Revised: August 26, 2020 S., Batut, P., Chaisson, M., and Gingeras, T.R. (2013). STAR: ultrafast universal
Accepted: October 2, 2020 RNA-seq aligner. Bioinformatics 29, 15–21.
Published: October 8, 2020 Finkel, Y., Mizrahi, O., Nachshon, A., Weingarten-Gabbay, S., Yahalom-Ro-
nen, Y., Tamir, H., Achdout, H., Melamed, S., Weiss, S., Israely, T., et al.
REFERENCES (2020). The coding capacity of SARS-CoV-2. Nature. Published online
September 9, 2020. https://doi.org/10.1038/s41586-020-2739-1.
Akopian, D., Shen, K., Zhang, X., and Shan, S.O. (2013). Signal recognition Gaglia, M.M., Covarrubias, S., Wong, W., and Glaunsinger, B.A. (2012). A com-
particle: an essential protein-targeting machine. Annu. Rev. Biochem. 82, mon strategy for host RNA degradation by divergent viruses. J. Virol. 86,
693–721. 9527–9530.
Alberti, S., Saha, S., Woodruff, J.B., Franzmann, T.M., Wang, J., and Hyman, Graham, R.L., and Baric, R.S. (2010). Recombination, reservoirs, and the
A.A. (2018). A User’s Guide for Phase Separation Assays with Purified Pro- modular spike: mechanisms of coronavirus cross-species transmission.
teins. J. Mol. Biol. 430, 4806–4820. J. Virol. 84, 3134–3146.
Alexopoulou, L., Holt, A.C., Medzhitov, R., and Flavell, R.A. (2001). Recogni- Guttman, M., Amit, I., Garber, M., French, C., Lin, M.F., Feldser, D., Huarte, M.,
tion of double-stranded RNA and activation of NF-kappaB by Toll-like receptor Zuk, O., Carey, B.W., Cassady, J.P., et al. (2009). Chromatin signature reveals
3. Nature 413, 732–738. over a thousand highly conserved large non-coding RNAs in mammals. Nature
Ameismeier, M., Cheng, J., Berninghausen, O., and Beckmann, R. (2018). 458, 223–227.
Visualizing late states of human 40S ribosomal subunit maturation. Nature Guttman, M., Garber, M., Levin, J.Z., Donaghey, J., Robinson, J., Adiconis, X.,
558, 249–253. Fan, L., Koziol, M.J., Gnirke, A., Nusbaum, C., et al. (2010). Ab initio

Cell 183, 1325–1339, November 25, 2020 1337


ll
OPEN ACCESS Article
reconstruction of cell type-specific transcriptomes in mouse reveals the Liberzon, A., Birger, C., Thorvaldsdóttir, H., Ghandi, M., Mesirov, J.P., and
conserved multi-exonic structure of lincRNAs. Nat. Biotechnol. 28, 503–510. Tamayo, P. (2015). The Molecular Signatures Database (MSigDB) hallmark
Hadjadj, J., Yatim, N., Barnabei, L., Corneau, A., Boussier, J., Smith, N., Péré, gene set collection. Cell Syst. 1, 417–425.
H., Charbit, B., Bondet, V., Chenevier-Gobeaux, C., et al. (2020). Impaired type Lloyd, R.E. (2006). Translational control by viral proteinases. Virus Res.
I interferon activity and inflammatory responses in severe COVID-19 patients. 119, 76–88.
Science, eabc6027. Lokugamage, K.G., Narayanan, K., Nakagawa, K., Terasaki, K., Ramirez, S.I.,
Hansen, T.H., and Bouvier, M. (2009). MHC class I antigen presentation: Tseng, C.-T.K., and Makino, S. (2015). Middle East Respiratory Syndrome Co-
learning from viral evasion strategies. Nat. Rev. Immunol. 9, 503–513. ronavirus nsp1 Inhibits Host Gene Expression by Selectively Targeting mRNAs
Transcribed in the Nucleus while Sparing mRNAs of Cytoplasmic Origin.
Hermann, T. (2016). Small molecules targeting viral RNA. Wiley Interdiscip.
J. Virol. 89, 10970–10981.
Rev. RNA 7, 726–743.
Lokugamage, K.G., Hage, A., de Vries, M., Valero-Jimenez, A.M., Schinde-
Holtkamp, W., Lee, S., Bornemann, T., Senyushkina, T., Rodnina, M.V., and
wolf, C., Dittmann, M., Rajsbaum, R., and Menachery, V.D. (2020). Type I inter-
Wintermeyer, W. (2012). Dynamic switch of the signal recognition particle
feron susceptibility distinguishes SARS-CoV-2 from SARS-CoV. J. Virol.,
from scanning to targeting. Nat. Struct. Mol. Biol. 19, 1332–1337.
JVI.01410-20.
Hong, V., Presolski, S.I., Ma, C., and Finn, M.G. (2009). Analysis and optimiza- Los, G.V., Encell, L.P., McDougall, M.G., Hartzell, D.D., Karassina, N., Zim-
tion of copper-catalyzed azide-alkyne cycloaddition for bioconjugation. An- prich, C., Wood, M.G., Learish, R., Ohana, R.F., Urh, M., et al. (2008). HaloTag:
gew. Chem. Int. 48, 9879–9883. a novel protein labeling technology for cell imaging and protein analysis. ACS
Izon, D.J., Punt, J.A., Xu, L., Karnell, F.G., Allman, D., Myung, P.S., Boerth, Chem. Biol. 3, 373–382.
N.J., Pui, J.C., Koretzky, G.A., and Pear, W.S. (2001). Notch1 regulates matu- Lungu, G.F., Stoica, G., and Wong, P.K.Y. (2008). Down-regulation of Jab1,
ration of CD4+ and CD8+ thymocytes by modulating TCR signal strength. Im- HIF-1alpha, and VEGF by Moloney murine leukemia virus-ts1 infection: a
munity 14, 253–264. possible cause of neurodegeneration. J. Neurovirol. 14, 239–251.
Jao, C.Y., and Salic, A. (2008). Exploring RNA transcription and turnover in vivo Maier, H.J., Neuman, B.W., Bickerton, E., Keep, S.M., Alrashedi, H., Hall, R.,
by using click chemistry. Proc. Natl. Acad. Sci. USA 105, 15779–15784. and Britton, P. (2016). Extensive coronavirus-induced membrane rearrange-
Jensen, S., and Thomsen, A.R. (2012). Sensing of RNA viruses: a review of ments are not a determinant of pathogenicity. Sci. Rep. 6, 27126.
innate immune receptors involved in recognizing RNA virus invasion. J. Virol. Majumdar, D., Frankiw, L., Burns, C., Garcia-Flores, Y., and Baltimore, D.
86, 2900–2910. (2018). Programmed Delayed Splicing: A Mechanism for Timed Inflammatory
Kamitani, W., Huang, C., Narayanan, K., Lokugamage, K.G., and Makino, S. Gene Expression. bioRxiv. https://doi.org/10.1101/443796.
(2009). A two-pronged strategy to suppress host protein synthesis by SARS Mayr, C., and Bartel, D.P. (2009). Widespread shortening of 3’UTRs by alterna-
coronavirus Nsp1 protein. Nat. Struct. Mol. Biol. 16, 1134–1140. tive cleavage and polyadenylation activates oncogenes in cancer cells. Cell
Kato, H., Takeuchi, O., Sato, S., Yoneyama, M., Yamamoto, M., Matsui, K., 138, 673–684.
Uematsu, S., Jung, A., Kawai, T., Ishii, K.J., et al. (2006). Differential roles of Miller, D.M., Rahill, B.M., Boss, J.M., Lairmore, M.D., Durbin, J.E., Waldman,
MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature 441, J.W., and Sedmak, D.D. (1998). Human cytomegalovirus inhibits major histo-
101–105. compatibility complex class II expression by disruption of the Jak/Stat
Kempf, B.J., and Barton, D.J. (2008). Poliovirus 2A(Pro) increases viral mRNA pathway. J. Exp. Med. 187, 675–683.
and polysome stability coordinately in time with cleavage of eIF4G. J. Virol. 82, Nakagawa, K., Narayanan, K., Wada, M., and Makino, S. (2018). Inhibition of
5847–5859. Stress Granule Formation by Middle East Respiratory Syndrome Coronavirus
4a Accessory Protein Facilitates Viral Translation, Leading to Efficient Virus
Kim, S.-I., Oceguera-Yanez, F., Sakurai, C., Nakagawa, M., Yamanaka, S., and
Replication. J. Virol. 92, e00902-18.
Woltjen, K. (2016). Inducible transgene expression in human iPS cells using
versatile all-in-one piggybac transposons. Methods Mol. Biol. 1357, 111–131. Narayanan, K., Huang, C., Lokugamage, K., Kamitani, W., Ikegami, T., Tseng,
C.-T.K., and Makino, S. (2008). Severe acute respiratory syndrome coronavi-
Kim, D.-K., Knapp, J.J., Kuang, D., and Cassonnet, P. (2020a). A flexible
rus nsp1 suppresses host gene expression, including that of type I interferon,
genome-scale resource of SARS-CoV-2 coding sequence clones. Preprints.
in infected cells. J. Virol. 82, 4471–4479.
https://doi.org/10.20944/preprints202004.0009.v1.
Narayanan, K., Ramirez, S.I., Lokugamage, K.G., and Makino, S. (2015). Coro-
Kim, D., Lee, J.Y., Yang, J.S., Kim, J.W., Kim, V.N., and Chang, H. (2020b). The
navirus nonstructural protein 1: Common and distinct functions in the regula-
Architecture of SARS-CoV-2 Transcriptome. Cell 181, 914–921.e10.
tion of host and viral gene expression. Virus Res. 202, 89–100.
Kochs, G., Garcı́a-Sastre, A., and Martı́nez-Sobrido, L. (2007). Multiple anti- Naus, J.I. (1982). Approximations for Distributions of Scan Statistics. J. Am.
interferon actions of the influenza A virus NS1 protein. J. Virol. 81, 7011–7021. Stat. Assoc. 77, 177.
König, J., Zarnack, K., Rot, G., Curk, T., Kayikci, M., Zupan, B., Turner, D.J., Ogle, J.M., Brodersen, D.E., Clemons, J., Tarry, M.J., Carter, A.P., and Ram-
Luscombe, N.M., and Ule, J. (2010). iCLIP reveals the function of hnRNP par- akrishnan, V. (2001). Recognition of cognate transfer RNA by the 30S ribo-
ticles in splicing at individual nucleotide resolution. Nat. Struct. Mol. Biol. 17, somal subunit. Science 292, 897–902.
909–915.
Oh, W., Yang, M.R., Lee, E.W., Park, K.M., Pyo, S., Yang, J.S., Lee, H.W., and
Kuglstatter, A., Oubridge, C., and Nagai, K. (2002). Induced structural changes Song, J. (2006). Jab1 mediates cytoplasmic localization and degradation of
of 7SL RNA during the assembly of human signal recognition particle. Nat. West Nile virus capsid protein. J. Biol. Chem. 281, 30166–30174.
Struct. Biol. 9, 740–744.
Parker, M.S., Balasubramaniam, A., Sallee, F.R., and Parker, S.L. (2018). The
Kurosaki, T., Popp, M.W., and Maquat, L.E. (2019). Quality and quantity con- Expansion Segments of 28S Ribosomal RNA Extensively Match Human
trol of gene expression by nonsense-mediated mRNA decay. Nat. Rev. Mol. Messenger RNAs. Front. Genet. 9, 66.
Cell Biol. 20, 406–420.
Robinson, J.T., Thorvaldsdóttir, H., Winckler, W., Guttman, M., Lander, E.S.,
Langmead, B., and Salzberg, S.L. (2012). Fast gapped-read alignment with Getz, G., and Mesirov, J.P. (2011). Integrative genomics viewer. Nat. Bio-
Bowtie 2. Nature methods 9, 357. technol. 29, 24–26.
Liaud, N., Horlbeck, M.A., Gilbert, L.A., Gjoni, K., Weissman, J.S., and Cate, Rodriguez, J.M., Maietta, P., Ezkurdia, I., Pietrelli, A., Wesselink, J.J., Lopez,
J.H.D. (2019). Cellular response to small molecules that selectively stall protein G., Valencia, A., and Tress, M.L. (2013). APPRIS: annotation of principal and
synthesis by the ribosome. PLoS Genet. 15, e1008057. alternative splice isoforms. Nucleic Acids Res. 41, D110–D117.

1338 Cell 183, 1325–1339, November 25, 2020


ll
Article OPEN ACCESS

Rozewicki, J., Li, S., Mar Amada, K., Standley, D.M., and Katoh, K. (2019). Vermeire, K., Bell, T.W., Van Puyenbroeck, V., Giraut, A., Noppen, S., Liekens,
MAFFT-DASH: integrated protein sequence and structural alignment. Nucleic S., Schols, D., Hartmann, E., Kalies, K.U., and Marsh, M. (2014). Signal pep-
Acids Research 47, W5–W10. tide-binding drug as a selective inhibitor of co-translational protein transloca-
Schmidt, E.K., Clavarino, G., Ceppi, M., and Pierre, P. (2009). SUnSET, a tion. PLoS Biol. 12, e1002011.
nonradioactive method to monitor protein synthesis. Nat. Methods 6, Voorhees, R.M., and Hegde, R.S. (2015). Structures of the scanning and
275–277. engaged states of the mammalian SRP-ribosome complex. eLife 4, 1–21.
Séraphin, B., Kretzner, L., and Rosbash, M. (1988). A U1 snRNA:pre-mRNA Wang, H., Li, X., Li, T., Zhang, S., Wang, L., Wu, X., and Liu, J. (2020). The ge-
base pairing interaction is required early in yeast spliceosome assembly but netic sequence, origin, and diagnosis of SARS-CoV-2. Eur. J. Clin. Microbiol.
does not uniquely define the 50 cleavage site. EMBO J. 7, 2533–2538. Infect. Dis. 39, 1629–1635.
Simonetti, A., Guca, E., Bochler, A., Kuhn, L., and Hashem, Y. (2020). Struc- Webb, B., and Sali, A. (2016). Comparative protein structure modeling using
tural Insights into the Mammalian Late-Stage Initiation Complexes. Cell Rep. MODELLER. Curr. Protoc. Bioinformatics 54, 5.6.1–5.6.37.
31, 107497. Weston, C.R., and Davis, R.J. (2007). The JNK signal transduction pathway.
Sola, I., Mateos-Gomez, P.A., Almazan, F., Zuñiga, S., and Enjuanes, L. (2011). Curr. Opin. Cell Biol. 19, 142–149.
RNA-RNA and RNA-protein interactions in coronavirus replication and tran- , M., Lapouge, K., Juaire, K.D., Flemming, D., Pfeffer, S., and
Wild, K., Aleksic
scription. RNA Biol. 8, 237–248. Sinning, I. (2020). MetAP-like Ebp1 occupies the human ribosomal tunnel exit
Stetson, D.B., and Medzhitov, R. (2006). Type I interferons in host defense. Im- and recruits flexible rRNA expansion segments. Nat. Commun. 11, 776.
munity 25, 373–381. Wilkins, C., and Gale, M., Jr. (2010). Recognition of viruses by cytoplasmic
Sutton, G., Fry, E., Carter, L., Sainsbury, S., Walter, T., Nettleship, J., Berrow, sensors. Curr. Opin. Immunol. 22, 41–47.
N., Owens, R., Gilbert, R., Davidson, A., et al. (2004). The nsp9 replicase pro- Yang, J., Anishchenko, I., Park, H., Peng, Z., Ovchinnikov, S., and Baker, D.
tein of SARS-coronavirus, structure and functional insights. Structure 12, (2020). Improved protein structure prediction using predicted interresidue ori-
341–353. entations. Proc. Natl. Acad. Sci. USA 117, 1496–1503.
Tanaka, Y., Kanai, F., Ichimura, T., Tateishi, K., Asaoka, Y., Guleng, B., Jazag, Zhang, Y., Ma, C., Yuan, Y., Zhu, J., Li, N., Chen, C., Wu, S., Yu, L., Lei, J., and
A., Ohta, M., Imamura, J., Ikenoue, T., et al. (2006). The hepatitis B virus X pro- Gao, N. (2014). Structural basis for interaction of a cotranslational chaperone
tein enhances AP-1 activation through interaction with Jab1. Oncogene 25, with the eukaryotic ribosome. Nat. Struct. Mol. Biol. 21, 1042–1046.
633–642. Zhang, L., Wang, J., Muñoz-Moreno, R., Kim, M., Sakthivel, R., Mo, W., Shao,
Tanaka, T., Kamitani, W., DeDiego, M.L., Enjuanes, L., and Matsuura, Y. D., Anantharaman, A., Garcı́a-Sastre, A., Conrad, N.K., and Fontoura, B.M.A.
(2012). Severe acute respiratory syndrome coronavirus nsp1 facilitates effi- (2018). Influenza Virus NS1 Protein-RNA Interactome Reveals Intron Targeting.
cient propagation in cells through a specific translational shutoff of host J. Virol. 92, e01634-18.
mRNA. J. Virol. 86, 11128–11137. Zhang, X., Tan, Y., Ling, Y., Lu, G., Liu, F., Yi, Z., Jia, X., Wu, M., Shi, B., Xu, S.,
Thoms, M., Buschauer, R., Ameismeier, M., Koepke, L., Denk, T., Hirschen- et al. (2020). Viral and host factors related to the clinical outcome of COVID-19.
berger, M., Kratzat, H., Hayn, M., Mackens-Kiani, T., Cheng, J., et al. (2020). Nature 583, 437–440.
Structural basis for translational shutdown and immune evasion by the Nsp1 Zhou, Q., Chen, V., Shannon, C.P., Wei, X.-S., Xiang, X., Wang, X., Wang, Z.-
protein of SARS-CoV-2. Science, eabc8665. H., Tebbutt, S.J., Kollmann, T.R., and Fish, E.N. (2020). Interferon-a2b Treat-
Van Nostrand, E.L., Pratt, G.A., Shishkin, A.A., Gelboin-Burkhart, C., Fang, ment for COVID-19. Front. Immunol. 11, 1061.
M.Y., Sundararaman, B., Blue, S.M., Nguyen, T.B., Surka, C., Elkins, K., Zou, L., Ruan, F., Huang, M., Liang, L., Huang, H., Hong, Z., Yu, J., Kang, M.,
et al. (2016). Robust transcriptome-wide discovery of RNA-binding protein Song, Y., Xia, J., et al. (2020). SARS-CoV-2 viral load in upper respiratory spec-
binding sites with enhanced CLIP (eCLIP). Nat. Methods 13, 508–514. imens of infected patients. N. Engl. J. Med. 382, 1177–1179.

Cell 183, 1325–1339, November 25, 2020 1339

You might also like